Environmental Fate, Transport, and Transformation of Carbon

Oct 18, 2011 - Smith, B.; Wepasnick, K.; Schrote, K. E.; Cho, H. H.; Ball, W. P.; Fairbrother, D. H. Influence of Surface Oxides on the Colloidal Stab...
0 downloads 0 Views 2MB Size
Chapter 4

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

Environmental Fate, Transport, and Transformation of Carbon Nanoparticles Liwen Zhang and Qingguo Huang* Department of Crop and Soil Sciences, University of Georgia, Griffin, Georgia *E-mail: [email protected]

Mass production of carbon nanoparticles (CNPs) is rapidly growing, and their entry to the environment is inevitable. Such releases of CNPs may cause undesired/unforeseen risks to the environment and human/wildlife health. A scientific assessment of such risks requires a thorough understanding of the environmental behaviors of CNPs, such as their fate, transport and transformation. This chapter presents a review on the important processes that govern the environmental behaviors of CNPs in natural aquatic systems, including aggregation, sorption, transport in porous media, and biotic and abiotic transformations.

1. Introduction Carbon nanoparticles (CNPs) are a family of nano-sized molecules composed almost entirely of carbon, with fullerene C60 and carbon nanotubes (CNTs) being the two most common types. C60 is a hollow spherical molecule, about 1 nm in diameter, comprised of 60 sp2 carbon atoms. CNTs are made in two principal classes: single- and multi-walled. Single-walled carbon nanotubes (SWCNTs) are one-layered graphitic cylinders with 1 - 5 nm diameters (4); whereas multi-walled carbon nanotubes (MWCNTs) comprise 2-30 concentric graphitic cylinders with outer diameters commonly between 10-100 nm. The lengths of CNTs vary widely, ranging between 10 nm to more than 1 cm (4). Several techniques, e.g., arc discharge, laser ablation, and chemical vapor deposition (CVD), have been used in CNP synthesis (10). Theoretically, pristine CNPs are purely composed of carbon atoms; however, in practice, defects cannot be avoided during synthesis © 2011 American Chemical Society

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

and subsequent purification processes. These defects usually impart the CNPs with hydroxyl or carboxyl surface functional groups. In addition, two types of methods are often used to modify CNP surfaces for increased solubility and biocompatibility, including: 1) strong oxidative treatments to create hydrophilic surface functional groups, and 2) the use of amphiphilic polymers or surfactants to wrap around and solubilize CNPs. Given the probable widespread application of manufactured CNPs, large-scale environmental release is possible. The total production capacity for nanocarbon products, including SWCNTs, MWCNTs, fullerenes, graphene, carbon nanofiber and nanodiamonds increased from 996 metric tons in 2008 to more than 2190 tons in 2009, and 4065 tons in 2010. The production capacity is expected to exceed 12,300 tons in 2015, reaching a compound annual growth rate of 24.8%. The growth is chiefly driven by multi-walled carbon nanotubes. World production capacity for multi-wall carbon nanotubes exceeded 390 tons in 2008, 1,500 tons in 2009, and was expected to exceed 3,400 tons in 2010 (11). CNPs may enter the environment through incidental release during manufacturing, transport and product use, or through waste disposal and decomposition (12, 13). Understanding the environmental behavior of CNPs is of paramount importance for an accurate environmental risk assessment. This review focuses on the processes that govern the important environmental behaviors of CNPs in natural aquatic systems. Aggregation, a process that controls the distribution of CNPs between solid and aqueous phases, is first discussed below. Sorption of CNPs on solid phases, a process that influences both CNP phase distribution and transport, is followed. The studies on transport behaviors of CNPs in porous media are then reviewed. Finally, recent investigation on transformation of CNPs under biotic and abiotic conditions is summarized.

2. Aggregation Fullerene C60 has limited solubility in some organic solvents (14) and is almost insoluble in water (5). Similarly, pristine CNTs could sparsely disperse in some organic solvents but not in polar solvents (15). As nanoscale particles (i.e., at least one dimesion is below 100 nm), CNPs can undergo Brownian motion and thus remain suspended in water over certain time scales as long as the settling velocity is equal to, or less than, the Brownian displacement. On the other hand, CNPs have large specific surface areas which are hydrophobic in nature, and are thus easy to aggregate leading to their settling from water. Aggregation is therefore a crucial process governing the phase distribution of CNPs. We will first in the following subsection (2.1) discuss the fundamentals of colloid science in relation to CNP aggregation, and then in subsequent subsections review observations on the effects of various system conditions on CNP aggregation. 2.1. Colloidal Nature Nanoparticles are essentially colloids, although towards the lower end of the colloid size range (Figure 1). As such, CNPs dispersed in water form a colloidal 70

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

dispersion which differs from a true solution in nature. A colloidal dispersion refers to one phase (e.g., solid) homogeneously distributed in another phase (e.g., water) (6). Colloidal dispersions are dynamic non-equilibrium systems, and are often sensitive to physical or chemical disturbances, which result in the aggregation of particles (16). There are two major steps involved in aggregation: particle transport (collision) and attachment (6, 17). The first step can be originated from three fundamental processes: Brownian diffusion of particles leads to perikinetic aggregation, shear flow transport of particles at different velocities causes orthokinetic (shear) aggregation, and particles of different size or density undergo differential settling (Figure 2). After initial aggregation, particle-cluster and cluster–cluster aggregation processes also take place (6).

Figure 1. Size domains and typical representatives of natural colloids and nanoparticles. The vertical line represents the operationally defined cut-off given by filtration at 0.45 µm. Used with permission (3).

71 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

Figure 2. The three collision mechanisms and associated rate coefficients for the aggregation of 1 µm particles with particles of diameter dp; the temperature is 12°C, the particle density 2.6 g/ml, and the shear rate 35/s. The cartoons represent the processes of perikinetic, orthokinetic, or differential settling, respectively. Dotted arrow indicates the graph relating to each cartoon (process). Used with permission (6). (see color insert) Generally the Brownian motion applies to a particle range of 1 nm—1 µm. When the particle size is beyond this range, the particles (or aggregates) begin settling. The settling rate for a spherical or near-spherical particle is proportional to the square of the particle diameter. When applied to non-spherical particles like CNTs, equivalent diameter can be calculated in terms of equivalent settling velocity or diffusivity. It is suggested that CNTs shorter than 500 nm can be simulated by ellipsoids (18), and the radius for a sphere of equivalent diffusivity can be calculated by Equation 1.

where , a and b are the radii of major and minor axes of the ellipsoid, respectively. Z is a function of s (18). Nonetheless, the effect of CNT shape on the aggregation and settling is still largely unknown and under debate. 72 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

Pristine CNPs have a strong tendency to attach and aggregate when they collide with each other as a result of attractive van der Waals forces. Experiments indicated that raw MWCNTs settled more rapidly than carbon black and activated carbon particles. This may be attributable to the much greater aspect ratio of MWCNTs (1:1,000), allowing for multiple contact points between particles, greater entanglement, and increased van der Waals forces, leading to aggregates of increased mass (19). Introduction of negative surface functional groups during CNP purification or modification tends to mitigate aggregation. Different types of surface functional groups can be added to CNPs during various purification and dispersion processes (20). These purification processes and their outcome will be discussed in greater detail later in the transformation section in this chapter. Such functional groups are in general hydrophilic and acidic (e.g., carboxyl, hydroxyl and carbonyl groups) (7, 8, 15, 21). Carboxyl groups, in particular, have low pKa values (~3.5), and are thus dissociated around common aqueous conditions rendering negative charges on CNP surfaces, although the charge densities may vary depending on the CNP synthesis and purification procedures. Smith et al. (12) found that the zeta-potential is proportional to the amount of carboxyl groups on CNT surfaces (R2 = 0.89). It was also found that up to six electrons can be accommodated in the lowest unoccupied molecular orbital (LUMO) of C60 despite its resistance to oxidation, which opens the route to covalent surface addition (22). Even in pure water, extended aqueous exposure can cause the initially hydrophobic C60 to form water-stable aggregates with externally positioned polar functional groups (23). These amphiphilic fullerene derivatives contain polar functions, such as hydroxyl, carboxyl and amino groups (24). Regardless of preparation methods, the surface charge of fullerenes and their aggregates are usually negative as indicated by their negative electrophoretic mobility or zeta potential, although the mechanisms are not fully understood yet (25). One explanation is that the core hydrophobic C60 molecules are likely cloaked by a polar shell formed via localized surface hydrolysis (e.g., C60 + H2O ↔ C60(OH)- + H+) (26). It has been found in several studies that the attachment efficiencies of both CNTs and fullerenes can be fairly well modeled using the Derjaguin—Landau—Verwey—Overbeek (DLVO) theory that describes the interactions of charged spherical colloidal particles. 2.2. Effect of Cation Concentration and Valence It is believed that the charges on CNP surfaces lead to repulsion forces that prevent CNP aggregation, and the surface potential of CNPs is correlated to their stability in aqueous systems (5, 12, 25, 27). It thus follows that any factor that can influence CNP surface potential may result in changes in the stability of CNPs in water. Consistent results have been reached in different studies that cations present in solution facilitated CNP aggregation (5, 9, 25, 27, 28). The cause has been rationalized, based on DLVO theory, to be the compression of the electric double layers (EDL) around CNPs by solution cations. Although DLVO theory was derived for ideal spherical particles with evenly distributed surface charge and 73

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

other surface properties, the aggregation behavior of CNTs can still be reasonably well modeled by DLVO theory in many aspects despite of their cylindrical shapes (27). In contrast to CNTs, fullerene and fullerene aggregates are more sphere-like, which renders them more amenable to measurement techniques and modeling based on electrophoretic mobility (5, 27) or zeta-potential (25, 27). As the concentration of electrolytes increased, the electrophoretic mobility (5, 27, 29) or zeta-potential (27, 30) of CNPs became less negative. For example, the zeta-potential of fullerene nanoparticles decreased linearly with the concentration of KCl (30). This suggests that the repulsive force between CNPs is reduced, and the attachment efficiency between CNPs would increase due to the decreasing energy barrier to aggregation. Once the concentration of electrolytes reaches or exceeds a certain value, i.e., the critical coagulation concentration (CCC), the particle surface charge becomes completely screened, thus eliminating the energy barrier to aggregation so that attractive forces between particles, e.g., van der Waals force, become dominant. The electrolytes commonly used in most studies cited above are NaCl, KCl, MgCl2 and CaCl2. Aggregation kinetics of the CNPs exhibited slow (reactionlimited or unfavorable) and fast (diffusion-limited or favorable) regimes in the presence of these cations, the intersection of which is the CCC (Figure 3) (5, 27, 28). The ability of divalent cations, e.g., Mg2+ or Ca2+, to induce aggregation is dramatically stronger than that of monovalent cations, e.g., Na+ or K+- (5, 9, 27). The CCC of Na+ for fullerene aggregation was 120 mM (5), while that of Ca2+ was 4.8 mM (5). According to the Schulze-Hardy rule, CCC is proportional to Z-6 for surfaces with high charge densities, or Z-2 for those with low charge densities (where Z is the counterion valence) (5, 9, 31). Observations with colloidal particles showed the CCC value dependence on Z ranged between Z-6 and Z-2 (Figure 4) (17). In a study with acid-treated MWCNTs, Smith et al. (27) found that the ratios of CCCs of MgCl2 and CaCl2 over that of NaCl were 2-5.7 and 2-6.3, respectively, very close to the theoretical value of 2-6 or 1/64. However, this theoretical ratio of 1/64 is not only determined by the valence of cations, but also the symmetry of electrolytes and the shape of CNPs. Researchers showed that for symmetric 2:2 electrolytes such as CaSO4 the ratio was 1/64, whereas for asymmetric 2:1 electrolytes, such as CaCl2, the ratio should be 1/42 (31, 32). Chen et al. (31) found the ratio of CCC values for colloidal particles with CaCl2 over NaCl was 1/40, consistent with the 1/42 prediction. Moreover, there were some experimental results not in good agreement with the 1/64 value, for example, 1/25 or 1/10 for CaCl2 (5) and 1/17 for MgCl2 (27). These discrepancies may be caused by different surface charge densities. While mono- and di-valent cations have large difference in CCCs, the difference between cations with the same valence is minor. The CCC values for an acid-treated MWCNTs (with carboxyl groups on surface) were 1.8 mM for MgCl2 and 1.2 mM for CaCl2 (27). Moreover, the type and valence of anions had little influence. The CCC values for the acid-treated MWCNTs were 93 mM Na+ in the form of NaCl and 98 mM Na+ for Na2SO4 (27). Although solution ionic conditions strongly influence the electric interactions between CNPs, the van der Waals forces are quite independent of solution conditions and CNP surface chemistry. Hamaker constant, a parameter delineating 74

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

van der Waals potential between two particles, does not change with solution conditions. The fitted Hamaker constants from two experiments investigating the aggregation of fullerene nanoparticles in aqueous medium were 6.7 × 10-21 J at pH 5.2 ± 0.1 (NaCl concentration from about 50 mM to 500 mM) (5) and 8.5 × 10-21 J at pH 5.5 (KCl concentration from about 20 mM to 1000 mM) (25). One additional point to be mentioned is that CNTs can be individually dispersed in water, at least for a large portion, while the water suspension of fullerenes is indeed their aggregations (nC60) instead of individual C60. The clusters have properties different from individual or bulk C60, which limits the application of individual C60 in aquatic systems. Comparing the CCC values of fullerenes, SWCNTs, and MWCNTs provides information on the relative aqeous stability of these CNPs. The CCC values were 20 or 37 mM Na+ for SWCNTs, 93 or 98 mM Na+ for MWCNTs, and 120 or 160 mM Na+ for fullerene (5, 24, 27, 33), indicating an increase of stability in this order. The CCC values of the divalent cation Ca2+ for SWCNT (~2 mM) (32) and MWCNT (1.2 mM) (27) were similar with each other, while that for fullerene was higher (6.1 mM) (24, 33).

2.3. Effect of pH CNP aggregation is also strongly influenced by pH, mainly because of the protonation/deprotonation of surface functional groups. Dissociation of surface functional groups contributes to surface charges. For example, the zeta-potential of fullerene nanoparticles became more negative when pH increased from 2 to 12, indicating more charges added (25), which led to smaller aggregate sizes and more stable suspension (28). Similar observations were made with CNTs (15). An increase in solution pH from 3 to 11 resulted in a substantial (over 2 orders of magnitude) decrease in MWCNT aggregation kinetics (27). Acid-treated MWCNTs were unstable at pH 0 but the stability increased when pH increased from 4 to 10 (27, 34). However, the electrophoretic mobility did not change much when pH was above 6 (27), indicating that electrophoretic mobility is not necessarily consistent with colloidal stability (12, 27).

75 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

Figure 3. (a) Attachment efficiencies of fullerene nanoparticles as a function of NaCl concentration at pH 5.2. The critical coagulation concentration (CCC) based on these data is 120 mM NaCl. (b) Attachment efficiencies of fullerene nanoparticles as a function of CaCl2 concentration at pH 5.2. The CCC based on these data is 4.8 mM CaCl2. The lines (used as eye guides) are extrapolated from the reaction-limited and diffusion limited regimes, and their intersections yield the respective CCC. Used with permission (5).

76 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

Figure 4. Double logarithmic plot of the critical coagulation concentrations (CCC) against the cation valance. The solid line has a slope of -6. Used with permission (9). 2.4. Effect of CNP Preparation Methods Different CNP synthesis approaches and the methods used to prepare CNP suspension will result in different CNP surface chemistries, and consequently different aggregation behavior. Several methods have been used to prepare fullerene water suspensions. One is the solvent exchange method comprising two steps: 1) dissolve fullerene in an organic solvent, e.g., tetrahydrofuran (THF) (20, 25, 35) or toluene (5, 20, 25, 35), and 2) introduce the mixture into water followed by removing the solvent through distillation (20, 35) or sonication (5, 20, 25, 35, 36). Another method is prolonged stir or ultrasonication of fullerenes in water (20, 25, 25, 25, 37). Dissimilarities were found between the nC60 produced by different methods, with respect to size, morphology, charge and hydrophobicity (20, 35, 37). For example, the relative hydrophobicity expressed as a partitioning coefficient to dodecane from water was 3.6% and 0.8% for nC60 prepared by solvent exchange using THF and extended mixing in water, respectively (20). Chen et al. (25) reported that the CCC value for fullerene stirred in water is 166 mM KCl, which is significantly greater than the CCC value (40 mM KCl) for fullerene suspension prepared by the solvent exchange method with toluene. These results indicate that a fullerene suspension prepared by solvent exchange seemed to be more stable than that prepared by mixing with water. The type of organic solvent used in the solvent exchange also made a difference; e.g., the nC60 clusters prepared in tetrahydrofuran had larger sizes than those prepared 77

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

in toluene (35). Even when the same stir/sonication method was used to prepare fullerene suspensions, the CCC values may still vary with specific procedure parameters (e.g. mixing time, C60-water ratio). Bouchard et al. (28) reported a CCC value for fullerene stirred in deionized distilled water for 5 months (100 mg C60 in 400 mL water) of 260 mM NaCl, while Chen et al. (25) reported that for a suspension prepared by stirring in deionized water for 40 days (1.22 g C60 in 1 L water) the CCC was 166 mM KCl. These phenomena can be attributed to the differences in the solvent properties. This led to different solvent-C60 interactions, the presence of different residual solvent in the nC60 structure, and thus different processes of the nC60 cluster formation (20, 38, 39). Solvent exchange is not commonly applied in preparing CNTs suspensions, because the solubility of CNTs is very low even in organic solvents. Methods widely used for preparing CNT suspensions include: 1) oxidization (e.g., acidtreatment, ozonation, etc.); 2) sonication; and 3) stirring in surfactant (or NOM) solutions. The first two methods create hydrophilic surface functional groups on CNTs, the effects of which will be discussed in section 5.1.1. CNTs with these hydrophilic groups tend to be more stable than pristine ones, because the more hydrophilic groups, such as hydroxyl and carboxyl groups, increase particle hydration and thus reduce the probability of particle-to-particle attachment during Brownian motion (19). Sonication tends to yield less stable CNT suspensions than the acid-treatment method, as indicated by comparison of their CCC values. For example, the CCC value was 20 mM NaCl for a sonicated suspension of SWCNTs, while that was 37 mM NaCl for the suspension of HNO3-treated SWCNTs (9, 33). The effects of surfactants and NOM on the stability of CNTs in water solutions are discussed in the following section. 2.5. Effect of Natural Organic Matter (NOM) Natural organic matter (NOM) ubiquitously exists in natural or engineered aquatic systems. It has been found in many studies that the presence of NOM significantly enhances the stability of CNPs in water (19, 20, 24, 28, 35, 37, 39) In the presence of Suwannee River NOM at the concentrations from 1 to 100 mg/L, CNTs and fullerenes were found to be more stable in aqueous phase (5, 19, 24, 27, 28, 33, 35, 39), with aggregation rates reduced and aqueous concentrations increased. Similar results were obtained in aggregation experiments using natural river water containing NOM (39, 40) and soluble soil humic substances (dissolved Aldrich humic acid at 150 mg/L and water-extractable Catlin soil humic substances at 300 mg/L) (39). To quantitatively investigate the effect of NOM, an experiment was carried out with various NOM concentrations and a fixed initial MWCNT concentration. It was found that, after 24 hours of settling, the concentration of MWCNTs remaining in water was linearly correlated with the NOM concentration varying from 0 to 100 mg/L (19, 39). In general, the concentration of MWCNT stably suspended in water is dependent on the amount of NOM adsorbed per unit mass of MWCNT (39). Similarly, NOM can cause disaggregation of nC60 crystals and aggregates under typical solution conditions of natural water (35). Moreover, microscopic and dynamic light scattering examination showed the NOM causes disaggregation, resulting in MWCNTs individually dispersed (39), and smaller 78

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

sizes of fullerene aggregates (35, 37, 39). These effects increased with increasing NOM concentration (35). It is believed that NOM may sorb on CNPs and then exert steric or electrostatic stabilization (5, 24, 28, 35, 39, 39, 41). NOM is in nature surface active, and because of this nature, their interactions with CNPs tend to take the form where the hydrophobic moieties of NOM associate with the CNP surface, likely through pi-pi or CH-pi stacking, while their hydrophilic moieties are exposed to the water (21, 39–42). Studies have shown that NOM association on CNTs was an exothermic equilibrium process, similar to spontaneous adsorbate-adsorbent interactions (39, 40), and followed pseudo-first-order rate kinetics (40). Hyung et al. (39) demonstrated the adsorption of NOM on CNT was proportional to the aromatic carbon content and molecular weight of the NOM. Pi-pi stacking is likely more powerful than CH-pi stacking. The stabilization effect of SDS, which is through CH-pi stacking, was substantially weaker than that of NOM which is through pi-pi stacking (43). The organic matter sorbed on CNPs imposes either steric and/or electrostatic stabilization, depending on their types, i.e., nonionic or ionic (41), thus preventing CNPs from attachment and aggregation (27, 28, 33, 41). The NOM that is sorbed on CNP surfaces has their hydrophilic portion extending into the solution phase. When two CNP particles approaching, the hydrophilic portion of the NOM interpenetrates and displaces water molecules, leading to steric stabilization (41). This stabilization effect is relatively inert to the change of ionic strength in the solution phase. For example, the stability of fullerenes in aqueous phase did not change significantly across a range of NaCl concentrations in the presence of NOM (24, 44). If the organic matter is ionic, both steric and electrostatic repulsion took place and the latter effect was influenced by the ionic strength of the solution (41). On the contrary, when adding divalent cations such as CaCl2, bridging effect or complex formation may occur and destabilize CNPs in the presence of NOM (24, 41, 44). The aggregation status of CNPs also influences the sorption of organic compounds. For example, theoretical calculations and nitrogen adsorption analysis results demonstrated that aggregation of CNTs led to a significant reduction in surface area (especially for SWCNTs), but a significant increase of pore volume (especially for MWCNTs) due to the interstices trapped in CNT aggregates (45). However, the adsorption of organic compounds on CNTs seemed to be controlled to a greater extent by the surface area rather than the pore volume in aqueous systems (45). Solution conditions also affect NOM sorption on CNPs. Higher ionic strength and lower pH both lead to NOM forming more coiled and compact structures (39, 46). This hampers the effect of NOM to stabilize CNTs via steric hindrance. Besides, ions reduce the charge potential of charged moieties on organic molecules. On the other hand, sorption of NOM increases as ionic strength increases or pH decreases, which tend to promote CNT stabilization (39, 42). Thus, the interplay of these two opposite effects determines how ionic strength or pH influences CNP stability in the presence of NOM. Generally, the net result of NOM is to assist CNP suspension. For example, humic acid and alginate (polysaccharide) exerts steric stabilization of CNPs with NaCl and 79

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

MgCl2 present in the solutions (28, 33, 39). In contrast, in the presence of CaCl2 the aggregation or deposition rates were larger with alginate on SWCNT (only at high Ca2+ concentrations) or C60 (even at low concentration of 0.3 mM) than the systems without alginate (28, 33, 39). Such enhanced aggregation was attributed to organic molecule bridging by Ca2+ (33, 39). Such bridging effect by Ca2+ was also observed in humic acid stabilized C60 nanoparticles (28). However, such bridging effect was not observed with other di-valent cations, e.g., Mg2+ .

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

2.6. Effect of Surface Functional Groups The speciation of surface functional groups is another influencing factor. For MWCNTs, Kennedy et al. (19) showed that the stabilizing ability of hydroxyl groups is greater than carboxyl groups, whereas Smith et al. (12) found that carboxyl groups are more influential than hydroxyl groups or carbonyl groups. It has been reported that the amount of surface oxygen-containing functional groups on CNTs correlated with their surface charge density (12), and the increased charge density tends to stabilize CNTs in aqueous phase. Similar to CNTs, the hydrophilic oxygen-containing surface functional groups on fullerenes also help to increase their stability in aqueous phase. For example, the CCC values for PCBM ([6, 6]-phenyl C61-butyric acid methyl ester) modified fullerenes were significantly higher than that of nC60 (28). When mixing with water, surface hydroxylation of the initially hydrophobic C60 molecules appeared to turn the nC60 clusters into hydrophilic, which helped stabilizing them in suspensions (21).

3. Sorption Solid phases (sediment or soil) ubiquitously exist in natural aquatic systems. The association of CNPs with these solid phases, or sorption, is another important process governing the partition of CNPs between water and solid phases. The extent to which C60 partitions to soil or sediment will influence its bioavailability and toxicity (39, 47), yet sorption of CNPs, especially CNTs, has not been extensively investigated. Sorption of fullerene by soils has been found to follow a linear isotherm with solid phase concentration proportional to aqueous phase concentration (39, 48). Soil organic matter (SOM) plays an important role in the sorption of nC60, and the sorption capacity strongly depends on the organic content of the soil (39, 48). Swelling of clay minerals also contributes to the sorption of C60 to soil (48). It was found that at low clay to organic carbon ratios, C60 sorption was dominated by SOM because much of the clay surface was coated by SOM (49). However, at higher clay to SOM ratios (fcm/foc > 20), the sorption of C60 by the swelling clay became influential (48). This result is consistent with the sorption of organic pollutants to soil. The sorption of C60 to SOM was found to depend on the SOM type and properties. If the organic matter is hydrophilic or surface active in nature, steric hindrance may take place and thus exert a stabilizing effect. For example, in a deposition study with silica as a solid phase in solutions containing NaCl, the attachment efficiencies between C60 and silica surface was mitigated when the silica surface was pre-coated with 80

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

dissolvable humic acid or alginate (28). This result implies that the negatively charged organic moieties or functional groups coated on soil minerals may reduce the sorption of fullerenes by soil. Studies related to CNT sorption on soil are limited and little is known about the interactions involved in the sorption. One type of soil organic matter (Canadian peat) has been found to sorb acid-treated MWCNTs from solutions containing cations (Na+), whereas, in the absence of cations, the sorption was not significant (50). This was attributed to that the cations caused a decrease in the surface charges of the soil organic material and CNTs, which facilitated interactions between them (50). An inorganic clay particle, kaolin, was found to improve MWCNT removal from aqueous phase (51), indicating a favorable association between kaolin and MWCNTs. The sorption of functionalized MWCNTs by soils followed a linear sorption isotherm pattern (52), whereas modifications of MWCNTs with polyethyleneimine (PEI) procedures to yield positive, negative, or neutral surface charges led to more non-linear sorption isotherm patterns (52). Soils also indirectly affect CNT stability in water. In a study investigating the interaction between clay minerals (kaolinite and montmorillonite) and MWCNT suspensions stabilized by surfactant (SDBS, CTAB, and TX100), clay minerals reduced the stability of MWCNTs in two ways: 1) competitive adsorption of surfactants thus reducing their stabilizing effect and 2) bridging between clay mineral and MWCNTs by surfactants (53). These effects depend on the properties of surfactants and the sorption capability of clay minerals. Additional research is needed to investigate the sorption effect of different soils or soil components under different solution conditions, and the sorption of SWCNTs.

4. Transport It is important to understand CNP transport through porous media in order to assess their potential to migrate in natural and engineered systems such as groundwater aquifers and water treatment filters. Most earlier studies of CNP mobility in porous media focused on model solid phases (quartz sand or glass beads) in packed columns (28, 54), with only several exceptions examining heterogeneous soil materials (55). 4.1. CNP Transport in Porous Media Previous studies suggest CNP mobility is governed by physicochemical deposition (filtration) and/or straining (54), which are determined by interactions among the CNPs (sorbate), the porous media (sorbent) and the solution (54). Heterogeneous solid phases, such as soils and wastewater sludge, comprise both organic and mineral components that have a number of potential sorption sites for CNPs (56). Organic matter contains negatively charged carboxyl and phenolic surface functional groups, positively charged sulfhydryl and amino surface functional groups, and regions of hydrophobicity generated by clusters of aromatic and aliphatic moieties (56, 57). While most soil mineral surfaces are hydroxylated and often carry negative charge due to isomorphic substitution, 81

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

both positive and negative charged sites can exist in metal oxides and along the edges of clays depending on the solution pH. These charged sites of soil as well as surface functional groups on CNPs render the electrostatic interaction one of the major mechanisms governing CNP transport in soil. Researches using model solid phases, i.e., glass beads or quartz sand, showed that the repulsion between the electrical double layers (EDLs) of CNPs and stationary phases with surface charges of the same sign resulted in stability and mobility. Thus, screening of the EDLs would lead to more deposition and retention of CNPs qualitatively consistent with the conventional colloid deposition theories (28, 54, 55). When electrolyte (NaCl) concentration increased from 1 mM to 10 mM, the C/C0 value (the relative effluent concentration) of nC60 clusters (168 nm in diameter) decreased from 0.71 to 0.33 (28). Similar results have been achieved by several other studies (28, 54, 55). CaCl2 seemed to be more capable in increasing nC60 and CNT retention in columns (54, 55). Under high ionic strength, e.g., ≥ 3.0 mM KCl, deposition (filtration) was the dominant process for CNP retention, while under low ionic strength, physical straining may also play a role in the capture of CNPs (54). Incomplete breakthrough of carboxyl functionalized SWCNTs in deionized water was observed (C/C0 = 0.90 ) with quartz sand packed columns (54). On the contrary, in glass beads or Ottawa sand packed columns, minimal nC60 retention occurred, and the breakthrough coincided with the nonreactive tracer (Br-) with deionized water as the mobile phase (54). This shows that the straining effect is more likely to take place in CNT transport. Jaisi et al. (54) concluded that the shape, particularly the very large aspect ratio of SWCNTs, and their highly aggregated state contribute to the retention of SWCNTs through enhanced straining. A study using a natural soil as the stationary phase showed that strong physical straining governed and prevented SWCNT transport through the media, which were collectively attributed to the shape and aggregation of SWCNTs, as well as the heterogeneity in soil particle size, porosity and permeability (55). This strong retention is insensitive to changes in the ionic strength to above 0.3 mM KCl or 0.1 mM CaCl2 (55). Hydrophobic interaction is another force, in addition to electrostatic, that may influence CNP transport. Fullerene nanoparticles can be functionalized to be more hydrophilic derivatives, i.e., fullerols. It was found that the mobility of fullerols (1.2 nm in diameter, monodispersed) was greater than the nC60 cluster (168 nm, monodispersed) (54). This may of course result from the difference in particle sizes, because smaller particles tend to be more mobile when all other properties are similar, but differences in hydrophilicity of the two types of particles may also play a role. It was found in a study with fullerene flowing through a column packed with spherical glass beads at low NaCl concentration (0.001 M) that the breakthrough curve did not monotonically increase with the injection flow (28). Instead, the affinity of porous media for nC60 increased after approximately one pore volume, followed by increased passage (28). This “affinity transition” was attributed to the initial association of nC60 to stationary phase created favorable sites for further loading of nC60 (28). However, the affinity did not continue to increase with more nC60 flowthrough but rather decreased, which calls for further investigation into 82

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

these phenomena. When NaCl concentration was high (e.g., 0.01 M or 0.1 M), this affinity transition was greatly reduced or eliminated and the affinity of nC60 towards porous media kept decreasing (28). Lecoanet et al. (54) suggested that this decreasing affinity of SWCNT, fullerol and nC60 under 0.01 M NaCl was caused by saturation or blocking of deposition sites within the porous media. Electrolytes in feed solution influence CNP release from porous media columns. In studies of deposition and transport of fullerene nanoparticles, the retention of CNPs by stationary phases (silica-coated quartz or quartz sands) were partly reversible under high pH of 12 or 10 (5, 54), and when the mobile phase had only a low concentration of monovalent salt (KCl) (54). Otherwise, at high monovalent salt concentration or in the presence of divalent salt (CaCl2), the deposition of the fullerenes was mostly irreversible. Introduction of deionized water resulted in a sharply declined breakthrough curve, indicating the fast release of nC60 from glass beads and quartz sand (54). The property of the stationary phase is another influential factor with regard to CNP transport in porous media columns. It is generally acknowledged that the stationary phases comprising finer size particles tend to have greater retention ability (54). Wang et al. (54) found finer Ottawa sand (100-140 mesh) can retain 95% of nC60 particles, much more than those by 40- to 50-mesh quartz sands. Under the same flow conditions (1.0 mM CaCl2, Darcy velocity 2.8 m/day), the retention of nC60 or MWCNTs in glass bead columns was substantially lower than in the quartz sand columns (54). Sectioned column tests showed that the nC60 retention by glass beads decreased with distance from the column inlet. In contrast, the retention by quartz sand was relatively constant through the entire column, suggesting that nC60 deposition approached a limiting capacity (54). Besides, in glass bead columns the nC60 retention can be completely recovered by deionized water extraction, whereas the retention by quartz sand was slightly resistant to water extraction (54). The pore water velocity also influences CNP transport in porous media columns, and it is generally inversely related to the retention of CNPs (54). Liu et al. (54) demonstrated that with greater pore water velocities (>4.0 m/day) MWCNT mobility was greater than that with 0.42 m/day velocity. Fullerenes exhibited similar breakthrough behaviors at a higher flow rate (40 mL/min or Darcy velocity of 0.14 cm/sec), regardless of differences in surface chemistry and sizes of the packing materials (54). In addition, the aforementioned affinity transition, in which the affinity between nC60 and stationary phase increased at the beginning and switched to decreasing, only occured at high velocity (28). However, the removal of fullerene-based nanoparticles was independent of the flow velocity under these conditions (e.g. 10 mM NaCl, pH 7), which suggested that the time scales for fullerene particle attachment or reorganization on the collector surface were greater than the time scale for them to transport to the collectors (54). Similar to the results of aggregation studies, humic-like substances largely reduced the retention, while the polysaccharide-based NOM, such as those produced by algae or bacteria, tended to favor deposition of nC60 (54). Different transport behaviors were observed with nC60 and SWCNTs even at the same column and flow conditions. In a quartz sand column, nC60 with diameter 83

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

of 168 nm had lower mobility than SWCNTs in a feed flow containing 10 mM NaCl (54); whereas in a soil column, nC60 with diameter of 51 nm displayed lower deposition rate and more effective transport than SWCNTs in a flow containing 0.1 to 100 mM KCl (55), indicating the importance of the CNP size to the mobility. The nC60 transport appeared to be more sensitive than SWCNTs to changes in the cation concentration from 0.03 to 100 mM KCl in the flow solutions (55).

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

4.2. Model Simulation Deposition is a crucial process governing the transport of CNPs and can be fitted well to models (54). Analogous to aggregation, deposition can be modeled as a sequence of particle transport to the immobile surface or “collector” described by a collector efficiency, η0, and followed by attachment described by an attachment efficiency, α (54). The theoretical single collector efficiency, η0, is composed of contact efficiencies due to interception (ηI), sedimentation (ηG), and diffusion (ηD). These efficiencies have been well modeled for spherical particles flowing through a system with spherical collectors, which can be applied in simulating spherical or near-spherical particles such as fullerene and fullerene clusters (58). However, this may not be readily applicable to CNTs because they are not spherical particles. Small-angle light scattering and ultra small-angle X-ray scattering showed that the morphology of MWCNTs in water were rod-like, and such rod-like morphology was not at the length-scale comparable to individual MWCNTs (from 1 nm to 50 µm), but seemed to be formed by networks of carbon "ropes" enmeshed with polyelectrolyte dispersants (59). There has not been any attempt to date to model the transport of such rope-like particles. Liu et al. (54) derived a relationship to model the deposition of MWCNTs based on their rod-like morphology to a spherical collector system. The collection efficiency was divided into two parts. The efficiency due to interception (ηI) by “end contact” was defined as:

84 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

The efficiency due to interception (ηI) by “side contact” was defined as:

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

where l is the length of MWCNTs, dp and dc are the diameters of the particles and collectors, respectively. The contact efficiency due to sedimentation is calculated by:

The contact efficiency due to diffusion is calculated by:

The overall collector efficiency is the sum of the three efficiencies described above:

where ρp is the MWCNT density, ρ is the fluid density, u is the fluid viscosity, T is the absolute temperature, and k is the Boltzmann constant. To calculate ηG and ηD requires a friction factor; however, this factor has not been developed for a cylindrical particle. Instead, the friction factor developed for a prolate ellipsoid has been employed (54). As mentioned above, deposition is the dominant process governing retention of CNTs only at high flow velocity and ionic strength. To compensate, this model was incorporated with a site-blocking term, which yielded good agreement with observed results in quartz sand column experiments (54). When the collector efficiency η0 is available, the attachment efficiency α can be estimated via Equation (6) (54): 85 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

where rC is the radius of a spherical collector, E is the porosity of the porous medium, L is the length of the porous medium, and C and C0 are the particle concentrations in the column effluent and influent, respectively. Based on the equations above, Liu et al. (54) estimated attachment efficiency factors for acid-treated MWCNTs passing through columns packed with quartz sand or glass beads with 8 mM Na+ under different flow rates (0.42, 4.0, 21 and 43 m/day). The resulting α values were relatively constant (~0.14) for all experimental conditions, but this value was more than one order of magnitude greater than the theoretical value of 0.009 calculated from DLVO theory (54). Such a discrepancy suggests that there are factors that may impact MWCNT deposition or transport processes that had not been accounted for in the modeling. Patch-wise surface charge heterogeneity of the sand grains is likely to contribute to such deviation from classical DLVO theory (54).

5. Transformation Possible CNP transformation in natural or engineered systems can change the properties of CNPs and consequently affect their mobility and bioavailability. Transformation of CNPs under natural conditions has not been fully investigated. There are, however, investigations regarding CNP reactions in chemistry and chemical engineering studies, and this information may suggest likely routes of CNP transformation in natural environments. In general, there are three types of transformations that can occur to CNPs: covalent reactions, biodegradation, and reactions of surface functional groups. 5.1. Covalent Reactions The graphene structures of CNPs, although inert in general, are still open to covalent reactions to certain extent, which is primarily driven by the enormous strain engendered by the curvature of CNTs and spherical geometry of fullerenes (7, 60). For an sp2-hybridized (trigonal) carbon atom, planarity is strongly preferred, described by a so-called pyramidalization angle of θp = 0° (Figure 5); whereas an sp3-hybridized (tetrahedral) carbon atom requires θp = 19.5° (60). According to the geometry of C60, all of the sp2-hybridized carbon atoms have θp = 11.6°, which is closer to the tetrahedral structure. Thus, the conversion of sp2- to sp3-hybridization can release the strain and mitigate the strain of the rest of the 59 atoms (60), which is consequently favorable to covalent addition (60, 61). The end caps of nanotubes, if not closed by the catalyst particle, tend to be composed of highly curved (and hence unstable) fullerene-like hemispheres that are much more reactive than the sidewalls (7, 62). The reactivity of these end caps is similar to fullerene, depending on the degree of pyramidalization. The 86

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

strain energy of pyramidalization is roughly proportional to (7). Thus, the smaller the diameter of fullerenes or fullerene end caps, the larger the curvature is and consequently the more reactive it is. The reactivity of CNT sidewalls comes from the curvature-induced pyramidalization, analogous to but weaker than fullerene. In addition, misalignment of the pi-orbitals between adjacent pairs of conjugated carbon atoms would also contribute to CNT side-wall reactivity (63), for which calculations of torsional strain energies in conjugated organic molecules has provided some theoretical support (7, 64).

Figure 5. Pyramidalization angle. Used with permission from Niyogi et al. (7). Oxidization is a common form of covalent reactions occurring to CNPs and has been studied the most. Certain treatments under harsh conditions can even destroy CNPs (65), however, these oxidation conditions are not likely to appear in surface earth processes or in manufacturing systems. Some oxidization processes are commonly employed in CNP purification to remove impurities, and these processes generally lead to reduction of CNP sizes and addition of oxygen functional groups to CNP surfaces, changes that tend to enhance CNP mobility and perhaps bioavailability and toxicity.

5.1.1. CNT Oxidization The unit backbone structure of CNT sidewalls, a six-numbered conjugated SP2 carbon ring, is relatively inert to oxidation. However, the sidewalls contain defect sites such as pentagon-heptagon pairs called Stone-Wales defects, sp3-hybridized defects, and vacancies in the nanotube lattice (62, 66). The end caps and the defects on sidewalls are expected to be sites more susceptible to oxidization (7, 21). Major types of oxidants include strong acids, e.g., concentrated HNO3 (21, 67), mixtures of concentrated HNO3/H2SO4 (20, 21), KMnO4/ H2SO4 (21, 68), 87 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

K2Cr2O7/ H2SO4 (68), OsO4 (7, 68), and H2O2/H2SO4 (21), as well as strong energy inputs such as ultrasonication (20, 64). However, these strong oxidative processes are unlikely to occur naturally in the environment. Oxidative reagents used in wastewater treatments, e.g., ozone (69), Fenton’s reagent (21, 70) and photophenton reagent (71), are also effective in oxidizing CNTs. Photooxidation is one process that can possibly occur to CNTs in the environment. An study has reported that reactive oxygen species (ROS), such as 1O2, O2•-, and •OH, were produced in carboxylated SWCNT solutions when exposed to the sunlight or lamp light within the solar spectrum, and these radicals oxidized CNTs and modified their surfaces (72). The common results of oxidation are to open the end cap and introduce oxygen-containing surface functional groups like carboxyl, hydroxyl, carbonyl, and ester to attach on either the ends or the sidewalls of CNTs (8, 15, 21, 71), although minor differences exist with various treatments. For example, Fenton’s reagent (Fe2+- H2O2) is effective in introducing both carboxyl and hydroxyl groups, while Photophenton and UV/H2O2 processes mostly produce hydroxyl groups (21). The strong oxidation processes were also found to disrupt the aromatic ring system of CNTs (21), for example, sonicated MWCNTs were shorter and exhibited a narrower length distribution (27). The consequence of these modifications is to increase the surface charge and thus stability in water. In addition to adding surface functional groups, the oxidative treatment of fullerene-like caps and graphene layers generated oxidized polycyclic aromatic substances, which were like fulvic acids and remained sorbed on MWCNT surfaces in acidic and neutral solutions (21). As mentioned before, these sorbed organic matters can also help to stabilize MWCNTs in water.

5.1.2. Fullerene Oxidation As discussed above, fullerene is generally more reactive than CNTs. The oxidization of fullerene does not require overly strong oxidants; instead, mild conditions can result in C60 oxidization. Prolonged mixing in water can cause negative charge and hydration on fullerene cluster surfaces (21, 37, 73). In dilute aqueous solution, the hydroxylated fullerenes, i.e. fullerenol, can be extensively mineralized by simulated solar radiation (74). This mineralization can reach up to 28% or approximately 47% (74), which is pH- and oxygen-dependent (74). The pH dependence of the direct and sensitized photoreactions is attributed to changes in intramolecular hemiketal formation in fullerenol (74). In contrast, the nC60 clusters formed in water are less reactive than fullerenol. Under 254 nm UV light and simulated or natural sunlight, mineralization of nC60 clusters was not observed, but oxygen-containing groups like epoxides and ethers were introduced (75). Oxygen is necessary for these oxidative transformations (74–76), and reactive oxidative species have been detected in these systems, including superoxide ions and singlet oxygen (1O2) (74).The resulting products tend to have a weaker antibacterial effect than the parent nC60 (75). A variety of covalent reactions have been designed as modification techniques to increase the solubility of CNPs in water or organic solvents. These reactions 88

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

include carbene chemistry (15, 77), nitrene addition (15, 77), hydrogenation via Birch reduction (15, 77, 78), fluorination (79), alkylation (80), arylation (81), and 1,3-dipolar cycloaddition (82). These original works have been summarized in the reviews by Niyogi et al. (7) and Banerjee et al. (8) for SWCNTs, and by Diederich et al. (83) for fullerenes.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

5.2. Biodegradation Recent studies indicate that carboxylated SWCNTs can be transformed via mediation by typical soil enzymes such as horseradish peroxidase (HRP), but such transformation did not seem to occur to pristine SWCNTs (2, 84). During such transformations, SWCNT lengths were shortened, carboxyl groups were added to SWCNT surfaces, and CO2 was produced (2). The products of the enzymatic degradation were identified as shown in Figure 6 (2). Recently, the same group has found that the neutrophil myeloperoxidase, a peroxidase generated inside human cells, can degrade SWCNTs and the resulting nanotubes did not generate an inflammatory response when aspirated into the lungs of mice (85). Using 13C-labeling, Schreiner et al. (1) found fullerol, the hydroxylized derivative of C60, can be mineralized to CO2 in the presence of white rot fungi after 32 weeks of incubation. Additionally, the fungi can incorporate minor amounts of the fullerol carbon into their lipid biomass, indicating the microbial utilization of fullerene derivatives (1). Since fullerol can be easily produced by mixing fullerene C60 with water, this biodegradation is thus quite probable when C60 enters natural systems. Figure 7 illustrates some potential environmental fates of fullerenes. 5.3. Reactions on Surface Functional Groups There have been a variety of studies in chemistry to design reactions targeting CNP surface functional groups, carboxyl in particular, to tether additional moieties, including small molecules, macromolecules or even other particles, to further modify the CNPs for various application purposes. Figure 8 presents some of such reactions to the surface functional groups on CNPs.

6. Environmental Implications Most studies on CNP aggregation focused on homoaggregation, i.e., aggregation among CNPs, whereas in natural aquatic systems heteroaggregation between CNPs and natural occurring colloids is more likely to dominate, because natural colloids would largely exceed the amount of CNPs (41). The collision rate for perikinetic aggregation and differential settling is lower when particles are of the same size; hence a monodispersed dispersion tends to be more stable than polydispersed dispersions (6). This implies that in natural water where other colloids or microbes are present, CNP aggregation tendency may be stronger, thus increasing the possibility of CNPs residing in the solid phase (soil/sediments). 89 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

Figure 6. Products identified by LC-MS for HRP-mediated degradation of SWCNTs, including oxidized PAHs such as benzaldehyde (1), 1,2-benzenediol (2), cinnamaldehyde (3), and diphenylacetic acid (4). Used with permission (2).

Figure 7. Overview of the potential environmental fates of fullerenes. Used with permission (1). Aggregation, sorption and transport of CNPs is governed by the interaction of a number of factors, including CNP surface chemistry, aquatic conditions such as ionic strength, pH, NOM concentration, and solid phase properties. Ionic strengths typical to many natural waters tend to favor deposition and thus reduce the potential exposure of CNPs. Without NOM present in water, the stable suspension of CNPs can be easily eliminated by divalent cations at low concentrations. However, NOM was found to counteract the effect of cations and dramatically stabilize CNPs at environmentally relevant concentrations (e.g., 5 mg/L). The stabilization effect of NOM is of paramount significance with regard to the potential mobility and exposure of CNPs in natural aquatic systems. Such stabilization effects enable CNPs to be transported through a longer distance and spread in a wider range. There are, however, studies showing that nC60 stabilized by dissolved humic substances lost the toxicity typically associated with nC60 when the humic acid concentrations were as low as 0.05 mg/L (39). The presence of NOM thus could have strong impacts on both the mobility and toxicity of CNPs, the two factors determining the potential environmental risks. Wastewater 90

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

treatments are an important defensive line protecting potable water. Researchers found that colloidal nC60 aggregates and MWCNTs can be efficiently removed by a series of coagulants (e.g., alum coagulant), and in the processes of flocculation, sedimentation and filtration, while the efficiency of removal was dependent on various parameters such as pH, alkalinity, NOM contents and coagulant dosage (51, 86).

Figure 8. Schematic of common functionalization routes used to derivatize SWCNTs at ends and defect sites. Used with permission (8). The transformation occurring to CNPs in natural environments tends to reduce nanoparticle sizes and add on hydrophilic groups. Such changes can lead to greater CNP mobility and perhaps greater bioavailability and toxicity as well. Templeton et al. (87) found smaller, more mobile fractions of SWCNTs were more toxic towards an estuarine copepod than the larger fractions. Lovern and Klaper (88) found a similar inverse relationship between the aggregate particle size and toxicity in Daphnia magna exposed to fullerenes. However, after phototransformation, the 91 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

toxicity of nC60 derivatives seemed to be less than the pristine nC60 (75). It is thus important when assessing the long-term environmental risks of CNPs to take into account potential transformation of CNPs in environmental systems.

References

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

1.

Schreiner, K. M.; Filley, T. R.; Blanchette, R. A.; Bowen, B. B.; Bolskar, R. D.; Hockaday, W. C.; Masiello, C. A.; Raebiger, J. W. White-Rot Basidiomycete-Mediated Decomposition of C-60 Fullerol. Environ. Sci. Technol. 2009, 43 (9), 3162–3168. 2. Allen, B. L.; Kotchey, G. P.; Chen, Y. N.; Yanamala, N. V. K.; Klein-Seetharaman, J.; Kagan, V. E.; Star, A. Mechanistic investigations of horseradish peroxidase-catalyzed degradation of single-walled carbon nanotubes. J. Am. Chem. Soc. 2009, 131 (47), 17194–17205. 3. Christian, P.; Von der Kammer, F.; Baalousha, M.; Hofmann, T. Nanoparticles: structure, properties, preparation and behaviour in environmental media. Ecotoxicology 2008, 17 (5), 326–343. 4. Chamot, J. A. The longest carbon nanotubes you’ve ever seen; National Science Foundation Press Release: Washington, DC, 2007. 5. Chen, K. L.; Elimelech, M. Aggregation and deposition kinetics of fullerene (C60) nanoparticles. Langmuir 2006, 22, 10994–11001. 6. Handy, R. D.; von der Kammer, F.; Lead, J. R.; Hassellov, M.; Owen, R.; Crane, M. The ecotoxicology and chemistry of manufactured nanoparticles. Ecotoxicology 2008, 17 (4), 287–314. 7. Niyogi, S.; Hamon, M. A.; Hu, H.; Zhao, B.; Bhowmik, P.; Sen, R.; Itkis, M. E.; Haddon, R. C. Chemistry of single-walled carbon nanotubes. Accounts Chem. Res. 2002, 35 (12), 1105–1113. 8. Banerjee, S.; Hemraj-Benny, T.; Wong, S. S. Covalent surface chemistry of single-walled carbon nanotubes. Adv. Mater. 2005, 17 (1), 17–29. 9. Sano, M.; Okamura, J.; Shinkai, S. Colloidal nature of single-walled carbon nanotubes in electrolyte solution: The Schulze-Hardy rule. Langmuir 2001, 17 (22), 7172–7173. 10. (a) Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56−58. (b) Kroto, H. W.; Heath, J. R.; Obrien, S. C.; Curl, R. F.; Smalley, R. E. C-60 - The 3rd man - A citation classic commentary on C-60 buckminsterfullerene. CC/Eng. Tech. Appl. Sci. 1993, 36, 8−9. (c) Cassell, A. M.; Raymakers, J. A.; Kong, J.; Dai, H. J. Large scale CVD synthesis of single-walled carbon nanotubes. J. Phys. Chem. B 1999, 103, 6484. 11. Chung, H.; Son, Y.; Yoon, T. K.; Kim, S.; Kim, W. The effect of multi-walled carbon nanotubes on soil microbial activity. Ecotoxicol. Environ. Saf. 2011, 74 (4), 569–575. 12. (a) Mueller, N. C.; Nowack, B. Exposure modeling of engineered nanoparticles in the environment. Environ. Sci. Technol. 2008, 42 (12), 4447−4453. (b) Panessa-Warren, B. J.; Maye, M. M.; Warren, J. B.; Crosson, K. M. Single walled carbon nanotube reactivity and cytotoxicity following extended aqueous exposure. Environ. Pollut. 2009, 157 (4), 1140−1151. 92

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

13.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

14.

15.

16.

17.

18.

19.

20.

21.

(c) Smith, B.; Wepasnick, K.; Schrote, K. E.; Cho, H. H.; Ball, W. P.; Fairbrother, D. H. Influence of Surface Oxides on the Colloidal Stability of Multi-Walled Carbon Nanotubes: A Structure-Property Relationship. Langmuir 2009, 25 (17), 9767−9776. Oliver, J. Carbon nanotubes: Technologies and commercial prospects; NAN024C; BCC Research: March, 2007. (a) Scrivens, W. A.; Tour, J. M.; Creek, K. E.; Pirisi, L. Synthesis of C-14labeled C-60, its suspension in water, and its uptake by human keratinocytes. J. Am. Chem. Soc. 1994, 116 (10), 4517−4518. (b) Ruoff, R. S.; Tse, D. S.; Malhotra, R.; Lorents, D. C. Solubility of C-60 in a variety of solvents. J. Phys. Chem. 1993, 97 (13), 3379−3383. (a) Shieh, Y. T.; Liu, G. L.; Wu, H. H.; Lee, C. C. Effects of polarity and pH on the solubility of acid-treated carbon nanotubes in different media. Carbon 2007, 45 (9), 1880−1890. (b) Chen, J.; Hamon, M. A.; Hu, H.; Chen, Y. S.; Rao, A. M.; Eklund, P. C.; Haddon, R. C. Solution properties of single-walled carbon nanotubes. Science 1998, 282 (5386), 95−98. Hassellov, M.; Readman, J. W.; Ranville, J. F.; Tiede, K. Nanoparticle analysis and characterization methodologies in environmental risk assessment of engineered nanoparticles. Ecotoxicology 2008, 17 (5), 344–361. Elimelech, M.; Gregory, J.; Jia, X.; Williams, R. Particle deposition and aggregation: measurement, modelling and simulation. In Colloid and Surface Engineering Series; Butterworth-Heinemann: Oxford, 1995. Phelan, F. R.; Bauer, B. J. Comparison of steric effects in the modeling of spheres and rodlike particles in field-flow fractionation. Chem. Eng. Sci. 2009, 64 (8), 1747–1758. Kennedy, A. J.; Hull, M. S.; Steevens, J. A.; Dontsova, K. M.; Chappell, M. A.; Gunter, J. C.; Weiss, C. A. Factors influencing the partitioning and toxicity of nanotubes in the aquatic environment. Environ. Toxicol. Chem. 2008, 27 (9), 1932–1941. (a) Schierz, A.; Zanker, H. Aqueous suspensions of carbon nanotubes: Surface oxidation, colloidal stability and uranium sorption. Environ. Pollut. 2009, 157 (4), 1088−1094. (b) Park, T. J.; Banerjee, S.; Hemraj-Benny, T.; Wong, S. S. Purification strategies and purity visualization techniques for single-walled carbon nanotubes. J. Mater. Chem. 2006, 16 (2), 141−154. (c) Lin, Y.; Taylor, S.; Li, H. P.; Fernando, K. A. S.; Qu, L. W.; Wang, W.; Gu, L. R.; Zhou, B.; Sun, Y. P. Advances toward bioapplications of carbon nanotubes. J. Mater. Chem. 2004, 14 (4), 527−541. (d) Park, H. J.; Park, M.; Chang, J. Y.; Lee, H. The effect of pre-treatment methods on morphology and size distribution of multi-walled carbon nanotubes. Nanotechnology 2008, 19 (33), Article #335702. (e) Brant, J. A.; Labille, J.; Bottero, J. Y.; Wiesner, M. R. Characterizing the impact of preparation method on fullerene cluster structure and chemistry. Langmuir 2006, 22, 3878. (a) Kuznetsova, A.; Popova, I.; Yates, J. T.; Bronikowski, M. J.; Huffman, C. B.; Liu, J.; Smalley, R. E.; Hwu, H. H.; Chen, J. G. G. Oxygen-containing functional groups on single-wall carbon nanotubes: NEXAFS and vibrational spectroscopic studies. J. Am. Chem. Soc. 2001, 123 (43), 10699−10704. 93

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

22. 23.

24.

25.

(b) Ying, Y. M.; Saini, R. K.; Liang, F.; Sadana, A. K.; Billups, W. E. Functionalization of carbon nanotubes by free radicals. Org. Lett. 2003, 5 (9), 1471−1473. (c) Liu, J.; Rinzler, A. G.; Dai, H. J.; Hafner, J. H.; Bradley, R. K.; Boul, P. J.; Lu, A.; Iverson, T.; Shelimov, K.; Huffman, C. B.; Rodriguez-Macias, F.; Shon, Y. S.; Lee, T. R.; Colbert, D. T.; Smalley, R. E. Fullerene pipes. Science 1998, 280 (5367), 1253−1256. (d) Monthioux, M.; Smith, B. W.; Burteaux, B.; Claye, A.; Fischer, J. E.; Luzzi, D. E. Sensitivity of single-wall carbon nanotubes to chemical processing: an electron microscopy investigation. Carbon 2001, 39 (8), 1251−1272. (e) Wang, Z. W.; Shirley, M. D.; Meikle, S. T.; Whitby, R. L. D.; Mikhalovsky, S. V. The surface acidity of acid oxidised multi-walled carbon nanotubes and the influence of in-situ generated fulvic acids on their stability in aqueous dispersions. Carbon 2009, 47 (1), 73−79. (f) Li, W.; Bai, Y.; Zhang, Y. K.; Sun, M. L.; Cheng, R. M.; Xu, X. C.; Chen, Y. W.; Mo, Y. J. Effect of hydroxyl radical on the structure of multi-walled carbon nanotubes. Synth. Met. 2005, 155 (3), 509−515. (g) Hiura, H.; Ebbesen, T. W.; Tanigaki, K. Opening and purification of carbon nanotubes in high yields. Adv. Mater. 1995, 7 (3), 275−276. (h) Escobar, M.; Goyanes, S.; Corcuera, M. A.; Eceiza, A.; Mondragon, I.; Rubiolo, G. H.; Candal, R. J. Purification and functionalization of carbon nanotubes by classical and advanced oxidation processes. J. Nanosci. Nanotechnol. 2009, 9 (10), 6228−6233. (i) Labille, J.; Brant, J.; Villieras, F.; Pelletier, M.; Thill, A.; Masion, A.; Wiesner, M.; Rose, J.; Bottero, J. Y. Affinity of C-60 fullerenes with water. Fullerenes, Nanotubes, Carbon Nanostruct. 2006, 14, 307−314. (j) Labille, J.; Masion, A.; Ziarelli, F.; Rose, J.; Brant, J.; Villieras, F.; Pelletier, M.; Borschneck, D.; Wiesner, M. R.; Bottero, J.-Y. Hydration and dispersion of C60 in aqueous systems: The nature of water-fullerene interactions. Langmuir 2009, 25 (19), 11232−11235. Mauter, M. S.; Elimelech, M. Environmental applications of carbon-based nanomaterials. Environ. Sci. Technol. 2008, 42 (16), 5843–5859. Isaacson, C. W.; Kleber, M.; Field, J. A. Quantitative analysis of fullerene nanomaterials in environmental systems: A critical review. Environ. Sci. Technol. 2009, 43 (17), 6463–6474. (a) Wiesner, M. R.; Lowry, G. V.; Alvarez, P.; Dionysiou, D.; Biswas, P. Assessing the risks of manufactured nanomaterials. Environ. Sci. Technol. 2006, 40 (14), 4336−4345. (b) Chen, K. L.; Elimelech, M. Influence of humic acid on the aggregation kinetics of fullerene (C60) nanoparticles in monovalent and divalent electrolyte solutions. J. Colloid Interface Sci. 2007, 309, 126−134. (a) Brant, J.; Lecoanet, H.; Hotze, M.; Wiesner, M. Comparison of electrokinetic properties of colloidal fullerenes (n-C60) formed using two procedures. Environ. Sci. Technol. 2005, 39, 6343. (b) Deguchi, S.; Alargova, R. G.; Tsujii, K. Stable dispersions of fullerenes, C-60 and C-70, in water. Preparation and characterization. Langmuir 2001, 17 (19), 6013−6017. (c) McHedlov-Petrossyan, N. O.; Klochkov, V. K.; Andrievsky, G. V., Colloidal dispersions of fullerene C-60 in water: some properties and regularities of coagulation by electrolytes. J. Chem. Soc.-Faraday 94

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

26.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

27.

28.

29.

30.

31.

Trans. 1997, 93 (24), 4343−4346. (d) Chen, K. L.; Elimelech, M. Relating colloidal stability of fullerene (C60) nanoparticles to nanoparticle charge and electrokinetic properties. Environ. Sci. Technol. 2009, 43 (19), 7270−7276. Andrievsky, G. V.; Klochkov, V. K.; Bordyuh, A. B.; Dovbeshko, G. I. Comparative analysis of two aqueous-colloidal solutions of C-60 fullerene with help of FTIR reflectance and UV-Vis spectroscopy. Chem. Phys. Lett. 2002, 364 (1-2), 8–17. (a) Smith, B.; Wepasnick, K.; Schrote, K. E.; Bertele, A. H.; Ball, W. P.; O’Melia, C.; Fairbrother, D. H. Colloidal properties of aqueous suspensions of acid-treated, multi-walled carbon nanotubes. Environ. Sci. Technol. 2009, 43 (3), 819−825. (b) Saleh, N. B.; Pfefferle, L. D.; Elimelech, M. Aggregation kinetics of multiwalled carbon nanotubes in aquatic systems: Measurements and environmental implications. Environ. Sci. Technol. 2008, 42 (21), 7963−7969. (a) Brant, J.; Lecoanet, H.; Wiesner, M. R. Aggregation and deposition characteristics of fullerene nanoparticles in aqueous systems. J. Nanopart. Res. 2005, 7 (4-5), 545−553. (b) Chen, K. L.; Elimelech, M. Interaction of fullerene (C-60) nanoparticles with humic acid and alginate coated silica surfaces: Measurements, mechanisms, and environmental implications. Environ. Sci. Technol. 2008, 42 (20), 7607−7614. (c) Bouchard, D.; Ma, X.; Isaacson, C. Colloidal properties of aqueous fullerenes: Isoelectric points and aggregation kinetics of C60 and C60 derivatives. Environ. Sci. Technol. 2009, 43 (17), 6597–6603. (d) Chen, Q.; Saltiel, C.; Manickavasagam, S.; Schadler, L. S.; Siegel, R. W.; Yang, H. Aggregation behavior of single-walled carbon nanotubes in dilute aqueous suspension. J. Colloid Interface Sci. 2004, 280, 91−97. (e) Duncan, L. K.; Jinschek, J. R.; Vikesland, P. J. C-60 colloid formation in aqueous systems: Effects of preparation method on size, structure, and surface charge. Environ. Sci. Technol. 2008, 42, 173. (f) Chang, X.; Vikesland, P. J. Effects of carboxylic acids on nC60 aggregate formation. Environ. Pollut. 2009, 157 (4), 1072−1080. (g) Gao, J.; Youn, S.; Hovsepyan, A.; Llaneza, V. L.; Wang, Y.; Bitton, G.; Bonzongo, J.-C. J. Dispersion and toxicity of selected manufactured nanomaterials in natural river water samples: Effects of water chemical composition. Environ. Sci. Technol. 2009, 43, 3322−3328. Saleh, N. B.; Pfefferle, L. D.; Elimelech, M. Aggregation Kinetics of Multiwalled Carbon Nanotubes in Aquatic Systems: Measurements and Environmental Implications. Environ. Sci. Technol. 2008, 42 (21), 7963–7969. Chen, K. L.; Elimelech, M. Relating Colloidal Stability of Fullerene (C60) Nanoparticles to Nanoparticle Charge and Electrokinetic Properties. Environ. Sci. Technol. 2009, 43 (19), 7270–7276. (a) Hsu, J. P.; Kuo, Y. C. The critical coagulation concentration of counterions: Spherical particles in asymmetric electrolyte solutions. J. Colloid Interface Sci. 1997, 185 (2), 530−537. (b) Grolimund, D.; Elimelech, M.; Borkovec, M. Aggregation and deposition kinetics of mobile colloidal particles in natural porous media. Colloids Surf. A 2001, 191, 179−188. 95

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

32. Hsu, J.-P.; Kuo, Y.-C. The Critical Coagulation Concentration of Counterions: Spherical Particles in Asymmetric Electrolyte Solutions. J. Colloid. Interface Sci. 1997, 185 (2), 530–537. 33. Saleh, N. B.; Pfefferle, L. D.; Elimelech, M. Influence of biomacromolecules and humic acid on the aggregation kinetics of single-walled carbon nanotubes. Environ. Sci. Technol. 2010, 44, 2412–2418. 34. Shieh, Y. T.; Liu, G. L.; Wu, H. H.; Lee, C. C. Effects of polarity and pH on the solubility of acid-treated carbon nanotubes in different media. Carbon 2007, 43, 1880. 35. Xie, B.; Xu, Z. H.; Guo, W. H.; Li, Q. L. Impact of natural organic matter on the physicochemical properties of aqueous C-60 nanoparticles. Environ. Sci. Technol. 2008, 42 (8), 2853–2859. 36. Andrievsky, G. V.; Kosevich, M. V.; Vovk, O. M.; Shelkovsky, V. S.; Vashchenko, L. A. On the production of an aqueous colloidal solution of fullerenes. J. Chem. Soc., Chem. Commun. 1995 (12), 1281–1282. 37. (a) Ma, X.; Bouchard, D. Formation of aqueous suspensions of fullerenes. Environ. Sci. Technol. 2009, 43 (2), 330−336. (b) Duncan, L. K.; Jinschek, J. R.; Vikesland, P. J. C-60 colloid formation in aqueous systems: Effects of preparation method on size, structure, and surface charge. Environ. Sci. Technol. 2008, 42, 173−178. 38. Choudhury, N.; Pettitt, B. M. On the mechanism of hydrophobic association of nanoscopic solutes. J. Am. Chem. Soc. 2005, 127 (10), 3556–3567. 39. (a) Li, Q. L.; Xie, B.; Hwang, Y. S.; Xu, Y. J. Kinetics of C-60 fullerene dispersion in water enhanced by natural organic matter and sunlight. Environ. Sci. Technol. 2009, 43 (10), 3574−3579. (b) Chen, K. L.; Elimelech, M. Influence of humic acid on the aggregation kinetics of fullerene C60 nanoparticles in monovalent and divalent electrolyte solutions. J. Colloid Interface Sci. 2007, 309, 126. (c) Hu, X. L.; Liu, J. F.; Mayer, P.; Jiang, G. Impacts of some environmentally relevant parameters on the sorption of polycyclic aromatic hydrocarbons to aqueous suspensions of fullerene. Environ. Toxicol. Chem. 2008, 27 (9), 1868−1874. (d) Li, D.; Lyon, D. Y.; Li, Q.; Alvarez, P. J. J. Effect of soil sorption and aquatic natural organic matter on the antibacterial activity of a fullerene water suspension. Environ. Toxicol. Chem. 2008, 27 (9), 1888−1894. (e) Chappell, M. A.; George, A. J.; Dontsova, K. M.; Porter, B. E.; Price, C. L.; Zhou, P. H.; Morikawa, E.; Kennedy, A. J.; Steevens, J. A. Surfactive stabilization of multi-walled carbon nanotube dispersions with dissolved humic substances. Environ. Pollut. 2009, 157 (4), 1081−1087. (f) Hyung, H.; Fortner, J. D.; Hughes, J. B.; Kim, J. H. Natural organic matter stabilizes carbon nanotubes in the aqueous phase. Environ. Sci. Technol. 2007, 41 (1), 179−184. (g) Hyung, H.; Kim, J. H. Natural organic matter (NOM) adsorption to multi-walled carbon nanotubes: Effect of NOM characteristics and water quality parameters. Environ. Sci. Technol. 2008, 42 (12), 4416−4421. (h) Liu, Y. Q.; Gao, L.; Zheng, S.; Wang, Y.; Sun, J.; Kajiura, H.; Li, Y.; Noda, K. Debundling of single-walled carbon nanotubes by using natural polyelectrolytes. Nanotechnology 2007, 18 (36), Article #365702. (i) 96

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

40.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

41.

42.

43.

44.

45.

46. 47.

48. 49. 50.

51.

Terashima, M.; Nagao, S. Solubilization of [60]fullerene in water by aquatic humic substances. Chem. Lett. 2007, 36 (2), 302−303. Su, F. S.; Lu, C. S. Adsorption kinetics, thermodynamics and desorption of natural dissolved organic matter by multiwalled carbon nanotubes. J. Environ. Sci. Health, Part A: Toxic/Hazard. Subst. Environ. Eng. 2007, 42 (11), 1543–1552. (a) Chen, K. L.; Smith, B. A.; Ball, W. P.; Fairbrother, D. H. Assessing the colloidal properties of engineered nanoparticles in water: case studies from fullerene C-60 nanoparticles and carbon nanotubes. Environ. Chem. 2009, 7 (1), 10−27. (b) Lin, D. H.; Xing, B. S. Tannic acid adsorption and its role for stabilizing carbon nanotube suspensions. Environ. Sci. Technol. 2008, 42 (16), 5917−5923. (a) Yang, K.; Xing, B. S. Adsorption of fulvic acid by carbon nanotubes from water. Environ. Pollut. 2009, 157 (4), 1095−1100. (b) Pan, B.; Xing, B. S. Adsorption Mechanisms of Organic Chemicals on Carbon Nanotubes. Environ. Sci. Technol. 2008, 42 (24), 9005−9013. O’Driscoll, N. J.; Messier, T.; Robertson, M. D.; Murimboh, J. Suspension of multi-walled carbon nanotubes (CNTs) in freshwaters: Examining the effect of CNT size. Water, Air, Soil Pollut. 2010, 208 (1-4), 235–241. Chen, K. L.; Elimelech, M. Interaction of Fullerene (C(60)) Nanoparticles with Humic Acid and Alginate Coated Silica Surfaces: Measurements, Mechanisms, and Environmental Implications. Environ. Sci. Technol. 2008, 42 (20), 7607–7614. Zhang, S.; Shao, T.; Bekaroglu, S. S. K.; Karanfil, T. The Impacts of Aggregation and Surface Chemistry of Carbon Nanotubes on the Adsorption of Synthetic Organic Compounds. Environ. Sci. Technol. 2009. Ghosh, K.; Schnitzer, M. Macromolecular structures of humic substances. Soil Sci. 1980, 129 (5), 266–276. (a) Petersen, E. J.; Huang, Q. G.; Weber, W. J., Jr. Bioaccumulation of radio-labeled carbon nanotubes by Eisenia foetida. Environ. Sci. Technol. 2008, 42 (8), 3090−3095. (b) Petersen, E. J.; Akkanen, J.; Kukkonen, J. V. K.; Weber, W. J. Biological uptake and depuration of carbon nanotubes by Daphnia magna. Environ. Sci. Technol. 2009, 43 (8), 2969−2975. (c) Petersen, E. J.; Huang, Q. G.; Weber, W. J., Ecological uptake and depuration of carbon nanotubes by Lumbriculus variegatus. Environ. Health Perspect. 2008, 116 (4), 496−500. Chen, C.-Y.; Jafvert, C. T. Sorption of buckminsterfullerene (C60) to saturated soils. Environ. Sci. Technol. 2009, 43, 7370–7375. Karickhoff, S. W. Organic pollutant sorption in aquatic systems. J. Hydraul. Eng.-ASCE 1984, 110 (6), 707–735. Zhang, L.; Petersen, E. J.; Huang, Q. Phase distribution of 14C-labeled multiwalled carbon nanotubes in aqueous systems containing model solids: Peat. Environ. Sci. Technol. 2011, 45 (4), 1356–1362. Holbrook, R. D.; Kline, C. N.; Filliben, J. J. Impact of source water quality on multiwall carbon nanotube coagulation. Environ. Sci. Technol. 2010, 44 (4), 1386–1391. 97

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

52. Petersen, E. J. P. E. J.; Pinto, R. A.; Zhang, L. W.; Huang, Q. G.; Landrum, P. F.; Weber, W. J. Effects of polyethyleneimine-mediated functionalization of multi-walled carbon nanotubes on earthworm bioaccumulation and sorption by soils. Environ. Sci. Technol. 2011, 45 (8), 3718–3724. 53. Han, Z.; Zhang, F.; Lin, D.; Xing, B. Clay minerals affect the stability of surfactant-facilitated carbon nanotube suspensions. Environ. Sci. Technol. 2008, 42 (18), 6869–6875. 54. (a) Wang, Y. G.; Li, Y. S.; Pennell, K. D., Influence of electrolyte species and concentration on the aggregation and transport of fullerene nanoparticles in quartz sands. Environ. Toxicol. Chem. 2008, 27 (9), 1860−1867. (b) Li, Y. S.; Wang, Y. G.; Pennell, K. D.; Abriola, L. M., Investigation of the transport and deposition of fullerene (C60) nanoparticles in quartz sands under varying flow conditions. Environ. Sci. Technol. 2008, 42 (19), 7174−7180. (c) Lecoanet, H. F.; Bottero, J. Y.; Wiesner, M. R., Laboratory assessment of the mobility of nanomaterials in porous media. Environ. Sci. Technol. 2004, 38 (19), 5164−5169. (d) Liu, X.; O’Carroll, D. M.; Petersen, E. J.; Huang, Q.; Anderson, C. L., Mobility of multiwalled carbon nanotubes in porous media. Environ. Sci. Technol. 2009, 43, 8153−8158. (e) Jaisi, D. P.; Saleh, N. B.; Blake, R. E.; Elimelech, M., Transport and filtration of carbon nanotubes in porous media. Geochim. Cosmochim. Acta 2008, 72 (12), A422−A422. (f) Espinasse, B.; Hotze, E. M.; Wiesner, M. R., Transport and retention of colloidal aggregates of C-60 in porous media: Effects of organic macromolecules, ionic composition, and preparation method. Environ. Sci. Technol. 2007, 41 (21), 7396−7402. (g) Wang, Y. G.; Li, Y. S.; Fortner, J. D.; Hughes, J. B.; Abriola, L. M.; Pennell, K. D., Transport and retention of nanoscale C-60 aggregates in water-saturated porous media. Environ. Sci. Technol. 2008, 42 (10), 3588−3594. (h) Jaisi, D. P.; Saleh, N. B.; Blake, R. E.; Elimelech, M., Transport of Single-Walled Carbon Nanotubes in Porous Media: Filtration Mechanisms and Reversibility. Environ. Sci. Technol. 2008, 42 (22), 8317-−8323. (i) Lecoanet, H. F.; Wiesner, M. R., Velocity effects on fullerene and oxide nanoparticle deposition in porous media. Environ. Sci. Technol. 2004, 38 (16), 4377−4382. 55. Jaisi, D. P.; Elimelech, M. Single-Walled Carbon Nanotubes Exhibit Limited Transport in Soil Columns. Environ. Sci. Technol. 2009, 43 (24), 9161–9166. 56. Sposito, G. The Chemistry of Soils; Oxford University Press: New York, 2008. 57. Sposito, G. The surface chemistry of natural particles; Oxford University Press: New York, 2004. 58. Tufenkji, N.; Elimelech, M. Correlation equation for predicting singlecollector efficiency in physicochemical filtration in saturated porous media. Environ. Sci. Technol. 2004, 38 (2), 529–536. 59. Schaefer, D.; Brown, J. M.; Anderson, D. P.; Zhao, J.; Chokalingam, K.; Tomlin, D.; Ilavsky, J. Structure and dispersion of carbon nanotubes. J. Appl. Crystallogr. 2003, 36, 553–557.

98 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

60. Haddon, R. C. Chemistry of the fullerenes - the manifestation of strain in a class of continuous aromatic molecules. Science 1993, 261 (5128), 1545–1550. 61. Taylor, R.; Walton, D. R. M. The chemistry of fullerenes. Nature 1993, 363 (6431), 685–693. 62. Hirsch, A. Functionalization of single-walled carbon nanotubes. Angew. Chem., Int. Ed. 2002, 41 (11), 1853–1859. 63. (a) Haddon, R. C. Measure of nonplanarity in conjugated organic molecules - which structurally characterized molecule displays the highest degree of pyramidalization? J. Am. Chem. Soc. 1990, 112 (9), 3385−3389. (b) Srivastava, D.; Brenner, D. W.; Schall, J. D.; Ausman, K. D.; Yu, M. F.; Ruoff, R. S. Predictions of enhanced chemical reactivity at regions of local conformational strain on carbon nanotubes: Kinky chemistry. J. Phys. Chem. B 1999, 103 (21), 4330−4337. 64. Suslick, K. S. The chemical effects of ultrasound. Sci. Am. 1989, 260 (2), 80–86. 65. Chiang, I. W.; Brinson, B. E.; Huang, A. Y.; Willis, P. A.; Bronikowski, M. J.; Margrave, J. L.; Smalley, R. E.; Hauge, R. H. Purification and characterization of single-wall carbon nanotubes (SWNTs) obtained from the gas-phase decomposition of CO (HiPco process). J. Phys. Chem. B 2001, 105 (35), 8297–8301. 66. Yao, N.; Lordi, V.; Ma, S. X. C.; Dujardin, E.; Krishnan, A.; Treacy, M. M. J.; Ebbesen, T. W. Structure and oxidation patterns of carbon nanotubes. J. Mater. Res. 1998, 13 (9), 2432–2437. 67. Hu, H.; Yu, A. P.; Kim, E.; Zhao, B.; Itkis, M. E.; Bekyarova, E.; Haddon, R. C. Influence of the zeta potential on the dispersability and purification of single-walled carbon nanotubes. J. Phys. Chem. B 2005, 109 (23), 11520–11524. 68. Hwang, K. C. Efficient cleavage of carbon graphene layers by oxidants. J. Chem. Soc., Chem. Commun. 1995 (2), 173–174. 69. (a) Banerjee, S.; Hemraj-Benny, T.; Balasubramanian, M.; Fischer, D. A.; Misewich, J. A.; Wong, S. S. Ozonized single-walled carbon nanotubes investigated using NEXAFS spectroscopy. Chem. Commun. 2004, (7), 772−773. (b) Hemraj-Benny, T.; Bandosz, T. J.; Wong, S. S. Effect of ozonolysis on the pore structure, surface chemistry, and bundling of single-walled carbon nanotubes. J. Colloid Interface Sci. 2008, 317 (2), 375−382. 70. Wang, Y. H.; Shan, H. W.; Hauge, R. H.; Pasquali, M.; Smalley, R. E. A highly selective, one-pot purification method for single-walled carbon nanotubes. J. Phys. Chem. B 2007, 111 (6), 1249–1252. 71. Fan, C. L.; Li, W.; Li, X.; Zhao, S. J.; Zhang, L.; Mo, Y. J.; Cheng, R. M. Efficient photo-assisted Fenton oxidation treatment of multi-walled carbon nanotubes. Chin. Sci. Bull. 2007, 52 (15), 2054–2062. 72. Chen, C. Y.; Jafvert, C. T. Photoreactivity of carboxylated single-walled carbon nanotubes in sunlight: Reactive oxygen species production in water. Environ. Sci. Technol. 2010, 44 (17), 6674–6679. 99

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

73. Brant, J. A.; Labille, J.; Bottero, J. Y.; Wiesner, M. R. Characterizing the impact of preparation method on fullerene cluster structure and chemistry. Langmuir 2006, 22 (8), 3878–3885. 74. Kong, L.; Tedrow, O.; Chan, Y. F.; Zepp, R. G. Light-initiated transformations of fullerenol in aqueous media. Environ. Sci. Technol. 2009, 43, 9155–9160. 75. Lee, J.; Cho, M.; Fortner, J. D.; Hughes, J. B.; Kim, J.-H. Transformation of aggregated C60 in the aqueous phase by UV irradiation. Environ. Sci. Technol. 2009, 43, 4878–4883. 76. Hou, W.-C.; Jafvert, C. T. Photochemistry of aqueous C60 clusters: Evidence of 1O2 formation and its role in mediating C60 phototransformation. Environ. Sci. Technol. 2009, 43, 5257–5262. 77. (a) Chen, Y.; Haddon, R. C.; Fang, S.; Rao, A. M.; Lee, W. H.; Dickey, E. C.; Grulke, E. A.; Pendergrass, J. C.; Chavan, A.; Haley, B. E.; Smalley, R. E. Chemical attachment of organic functional groups to single-walled carbon nanotube material. J. Mater. Res. 1998, 13 (9), 2423−2431. (b) Holzinger, M.; Vostrowsky, O.; Hirsch, A.; Hennrich, F.; Kappes, M.; Weiss, R.; Jellen, F. Sidewall functionalization of carbon nanotubes. Angew. Chem.-Int. Edit. 2001, 40 (21), 4002. 78. Pekker, S.; Salvetat, J. P.; Jakab, E.; Bonard, J. M.; Forro, L. Hydrogenation of carbon nanotubes and graphite in liquid ammonia. J. Phys. Chem. B 2001, 105 (33), 7938–7943. 79. (a) Mickelson, E. T.; Huffman, C. B.; Rinzler, A. G.; Smalley, R. E.; Hauge, R. H.; Margrave, J. L. Fluorination of single-wall carbon nanotubes. Chem. Phys. Lett. 1998, 296 (1-2), 188−194. (b) Mickelson, E. T.; Chiang, I. W.; Zimmerman, J. L.; Boul, P. J.; Lozano, J.; Liu, J.; Smalley, R. E.; Hauge, R. H.; Margrave, J. L. Solvation of fluorinated single-wall carbon nanotubes in alcohol solvents. J. Phys. Chem. B 1999, 103 (21), 4318−4322. 80. Boul, P. J.; Liu, J.; Mickelson, E. T.; Huffman, C. B.; Ericson, L. M.; Chiang, I. W.; Smith, K. A.; Colbert, D. T.; Hauge, R. H.; Margrave, J. L.; Smalley, R. E. Reversible sidewall functionalization of buckytubes. Chem. Phys. Lett. 1999, 310 (3-4), 367–372. 81. (a) Bahr, J. L.; Tour, J. M. Highly functionalized carbon nanotubes using in situ generated diazonium compounds. Chem. Mat. 2001, 13 (11), 3823−3824. (b) Bahr, J. L.; Yang, J. P.; Kosynkin, D. V.; Bronikowski, M. J.; Smalley, R. E.; Tour, J. M., Functionalization of carbon nanotubes by electrochemical reduction of aryl diazonium salts: A bucky paper electrode. J. Am. Chem. Soc. 2001, 123 (27), 6536−6542. 82. Georgakilas, V.; Kordatos, K.; Prato, M.; Guldi, D. M.; Holzinger, M.; Hirsch, A. Organic functionalization of carbon nanotubes. J. Am. Chem. Soc. 2002, 124 (5), 760–761. 83. Diederich, F.; Thilgen, C. Covalent fullerene chemistry. Science 1996, 271 (5247), 317–323. 84. Allen, B. L.; Kichambare, P. D.; Gou, P.; Vlasova, I.; Kapralov, A. A.; Konduru, N.; Kagan, V. E.; Star, A. Biodegradation of single-walled carbon nanotubes through enzymatic catalysis. Nano Lett. 2008, 8 (11), 3899–3903. 100

In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.

Downloaded by UNIV OF GUELPH LIBRARY on June 2, 2012 | http://pubs.acs.org Publication Date (Web): October 18, 2011 | doi: 10.1021/bk-2011-1079.ch004

85. Kagan, V. E.; Konduru, N. V.; Feng, W. H.; Allen, B. L.; Conroy, J.; Volkov, Y.; Vlasova, II; Belikova, N. A.; Yanamala, N.; Kapralov, A.; Tyurina, Y. Y.; Shi, J. W.; Kisin, E. R.; Murray, A. R.; Franks, J.; Stolz, D.; Gou, P. P.; Klein-Seetharaman, J.; Fadeel, B.; Star, A.; Shvedova, A. A. Carbon nanotubes degraded by neutrophil myeloperoxidase induce less pulmonary inflammation. Nat. Nanotechnol. 2010, 5 (5), 354–359. 86. Hyung, H.; Kim, J. H. Dispersion of C-60 in natural water and removal by conventional drinking water treatment processes. Water Res. 2009, 43 (9), 2463–2470. 87. Templeton, R. C.; Ferguson, P. L.; Washburn, K. M.; Scrivens, W. A.; Chandler, G. T. Life-cycle effects of single-walled carbon nanotubes (SWNTs) on an estuarine meiobenthic copepod. Environ. Sci. Technol. 2006, 40 (23), 7387–7393. 88. Lovern, S. B.; Klaper, R. Daphnia magna mortality when exposed to titanium dioxide and fullerene (C-60) nanoparticles. Environ. Toxicol. Chem. 2006, 25 (4), 1132–1137.

101 In Biotechnology and Nanotechnology Risk Assessment: Minding and Managing the Potential Threats around Us; Ripp, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2011.